Accurate Theoretical Description of the 1La and 1Lb Excited States in

Aug 23, 2012 - Young Choon Park , Florian Senn , Mykhaylo Krykunov , and Tom Ziegler ... Monte Carlo and Many Body Green's Function Theory Calculation...
1 downloads 0 Views 6MB Size
Subscriber access provided by University of Guelph Library

Article

An Accurate Theoretical Description of the 1La and 1Lb Excited States in Acenes using the all Order Constricted Variational Density Functional Theory Method and the Local Density Approximation. Mykhaylo Krykunov, Stefan Grimme, and Tom Ziegler J. Chem. Theory Comput., Just Accepted Manuscript • DOI: 10.1021/ct300372x • Publication Date (Web): 23 Aug 2012 Downloaded from http://pubs.acs.org on September 8, 2012

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of Chemical Theory and Computation is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

An  Accurate  Theoretical  Description  of  the  1La  and  1Lb  Excited   States  in  Acenes  using    the  all  Order  Constricted  Variational   Density  Functional  Theory  Method  and  the  Local  Density   Approximation.   Mykhaylo  Krykunov,*,a      Stefan  Grimme,  b  and  Tom  Ziegler  a   Department  of  Chemistry,  University  of  Calgary,  University  Drive  2500,  Calgary,   Alberta  ,  Canada  T2N  1N4.   Mulliken  Center  for  Theoretical  Chemistry,  Institut  für  Physikalische  und   Theoretische  Chemie  der  Universität  Bonn,  Beringstr.  4,D-­‐53115,  Germany       Abstract:   We   present   the   results   of   calculations   on   the   vertical   singlet   1La   and   1Lb   excitation   energies   in   acenes   within   time   dependent   density   functional   theory   (TDDFT),   second   order   constricted   variational   DFT   (CV(2)-­‐DFT)   and   all   order   constricted   variational   DFT   (CV(∞)-­‐DFT)   using   the   local   density   approximation   LDA(VWN).   For   the   linear   acenes   it   is   shown   that   the   application   of   the   Tamm-­‐ Dancoff   (TD)   approximation   to   TDDFT   (TDDFT-­‐TD)   substantially   improves   the   agreement   with   experiment   compared   to   pure   TDDFT.   This   improvement   leads   to   the   correct   ordering   of   the   1La   and   1Lb   excitation   energies   in   Naphthalene.   As   TDDFT-­‐TD  is  equivalent  to  the  second  order  CV(2)-­‐TD  method  one  might  hope  for   further  improvements  by  going  to  all  orders  in  CV(∞)-­‐TD.  Indeed,   for  linear  acenes   the  application  of  the  CV(∞)-­‐TD  method  brings  the  agreement  with  experiment  to   within   0.1   eV   for   both   types   of   excitations   using   the   simple   LDA   functional.   The   CV(∞)-­‐TD   method   based   on   LDA   is   also   shown   to   be   accurate   for   15   nonlinear   acenes  with  root  mean  square  deviations  of  0.24  eV  for  1La  and  0.17  eV  for  1Lb.     1  Introduction   A   number   of   π-­‐conjugated   systems   in   the   form   of   organic   dyes   have   recently   attracted  considerable  attention  as  possible  candidates  that  might  be  incorporated   into   dye-­‐sensitized   solar   cells.   To   devise   the   optimal   cell,   it   would   be   very   helpful   to   have   an   accurate   theoretical   method   that   could   search   for   good   candidates   among                                                                                                                   *  Corresponding  author:  E-­‐mail  [email protected]   aUniversity  of  Calgary   b  University  of  Bonn    

ACS Paragon Plus Environment

1  

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 29

the  many  possible  dyes  in  a  both  efficient  and  accurate  manner.  In  order  to  establish   such   a   method,   it   is   also   necessary   to   have   a   set   of   benchmarks   among   π   →   π∗   excitation   energies   in   extended   π-­‐conjugated   systems.   Fortunately,   the   singlet   1La   and   1Lb  excitations  in  linear  poly-­‐acenes  represent  such  a  set  of  benchmark  data  as   they  have  been  studied  extensively  both  experimentally  and  theoretically.  Here  the   acenes  consist  of  a  number  (nr)  of  fused  benzene  rings.   The   distinct   properties1   of   the   1La   and   1Lb   states   for   the   linear   acenes   have   already  been  described  in  the  literature.2-­‐3    Essentially  the   1La  (or   1B2u  when  the  x-­‐ axis   corresponds   to   the   long   molecular   axis   )   state   is   dominated   by   a   single   electron   transition  HOMO  →  LUMO  ,  while  the   1Lb  (or   1B3u)  state  results  from  a  combination   of  HOMO-­‐1  →  LUMO  and  HOMO  →  LUMO+1  transitions.  Further,  excitations  to   1La   (1B2u)   are   short   axis   polarized   with   high   intensity   whereas   the   transitions   to   1Lb   (1B3u)   are   long   axis   polarized   with   low   intensity,   Table   1.   The   relative   ordering   of   the  1La  and  1Lb  states  is  size  dependent.  Thus,  for  naphthalene  with  two  rings  the  1Lb   state  is  below  1La  in  energy  while  for  the  other  linear  acenes  the  ordering  is  opposite   although   the   splitting   is   very   small   for   anthracene   with   nr   =3.   Both   states   drop   in   energy  with  increasing  nr  but  the  drop  is  largest  for  1La  ,  see  Table  1  and  Figure  1.     It  was  previously  reported2,3    that  TDDFT4-­‐7    incorrectly  predicts    the  crossing   point  of  the   1La  and   1Lb  states  in  linear  poly-­‐acenes  with  both  semi-­‐local  and  hybrid   functionals.     Further,   the   computed   TDDFT   excitation   energies   for   1La   relative   to   1Lb   are  in  poor  agreement  with  experiment  throughout  the  range  nr  =2,6.8-­‐9      Promising   alternative  methods  are  the  DFT/MRCI  approach11-­‐12  and  TDDFT  schemes  based  on  

 

ACS Paragon Plus Environment

2  

Page 3 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

long-­‐range   corrected   (LRC)   functionals.13-­‐16   The   DFT/MRCI   approach   is   quite   accurate  but  not  a  “black  box”  method  like  the  TDDFT  scheme.  This  makes  it  difficult   to   use   DFT/MRCI   in   screening   type   applications.   The   LRC   methods   show   some   improvement   for   1La   but   not   for   1Lb.16     Further,   the   optimal   range   separation   parameter(s)   must   be   determined   for   each   molecule.   A   more   recent   promising   approach   is   double-­‐hybrid   density   functional   theory.20-­‐25     This   scheme   has   to   date   provided  the  most  balanced  description  of  the  linear  acenes.25  Unfortunately,  these   double   hybrids   are   limited   to   low   valence   excitations   due   to   the   inherent   perturbative  corrections.26     Some   of   us   have   recently   put   forward27   the   constricted   variational   density   functional  theory  (CV-­‐DFT)  as  a  possible  improvement  of  adiabatic  TDDFT.  The  new   theory  is  in  its  simplest  second  order  formulation  (CV(2)-­‐DFT)  identical  to  adiabatic   TDDFT   after   use   has   been   made   of   the   popular   Tamm-­‐Dancoff   approximation.28     However   the   method   can   be   extended   to   all   orders   in   the   variational   parameters   (CV(∞)-­‐DFT).   It   has   been   shown   in   a   recent   study27   that   the   π   →   π∗   transitions   in   conjugated   hydrocarbons   are   described   reasonably   well   within   the   CV(∞)-­‐DFT   scheme.   An   average   deviation   from   experiment   is   approximately   0.3   eV   for   both   semi-­‐local  and  hybrid  functionals.     We   shall   here   first   assess   the   ability   of   our   newly   developed   constricted   variational   density   functional   theory   (CV-­‐DFT)27   to   describe   the   excited   states   in   linear  poly-­‐acenes,  as  a  first  step  in  an  attempt  to  establish  an  inexpensive  method   that   ultimately   can   help   in   the   design   of   light-­‐harvesting   devices.   It   is   further   our  

 

ACS Paragon Plus Environment

3  

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 29

objective   to   demonstrate   that   quite   accurate   results   can   be   obtained   for   our   test   set   even   with   the   simple   LDA   functional   if   one   goes   beyond   the   simple   second   order   approximation   inherent   in   TDDFT   and   incorporate   terms   to   all   orders   in   the   variational   parameters,   as   it   is   done   in     CV(∞)-­‐DFT.   Our   work   is   a   continuation   of   a   previous  study  in  which  we  demonstrated  that  LDA  can  be  used  to  describe  charge   transfer   transitions   within   the   constricted   variational   DFT   method.   29   We   shall   finally  extend  the  scope  of  our  investigation  to  include  the   1La  and   1Lb  excitations  in   15  nonlinear  poly-­‐acenes.   2  Computational  Method  and  Computational  Details   2.1  Constricted  Variational  Density  Functional  theory    

In  CV(n)-­‐DFT,  we  carry  out  a  unitary  transformation  among  occupied  

{" i ;i = 1, occ}  and  virtual   {" a ; a = 1, vir }    ground  state  orbitals  

  !

!!! occ                                                   Y # # !vir "

$ ! ! $ ! m=' m $! ! $ ! !' U occ occ U &=e # &=# ( & = # occ &# & & # !vir & # # & # ' % " % " m=0 m! %" !vir % #" !vir

$ & &&        (1)   %

with   the   summation   over   m   in   (1)   truncated   at   m=n.   Thus   the   resulting   orbitals   will   be   orthonormal   to   order   n   in   U .   Here   " oc c  and   " vir  are   concatenated   column   vectors   consisting  of  the  sets   {" i ;i = 1, occ} and   {" a ; a = 1, vir }  of  occupied  and  virtual  ground  

! !" '  and   " ' are   concatenated   column   vectors   state   orbitals,   respectively,   whereas   oc c vir ! ! ' ;i = 1, occ}  and   {" ' ;a = 1, vir}  of  occupied  and  virtual   containing  the  resulting  sets   {" i a

! ! unitary   transformation   matrix   Y  is   in   eq   1   excited   state   orbitals,   respectively.   The   ! matrix   U  as     expressed  in  terms  o!f  a  skew  symmetric  

!  

! Plus Environment ACS Paragon

4  

Page 5 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

! ! U2 U m ! (U 2 )m (U 2 )m                                     Y = e = I +U +    (2)   +! = " =" +U" 2 m=0 m! m=0 2m! m=0 (2m +1)! U

Here     Uij = U ab = 0  where   “i,j”   refer   to   the   occupied   set   {" i ; i = 1, occ }  whereas   “a,b”   refer   to   {" a ; a = 1, vir } .   Further,   U ai  are   the   variational   mixing   matrix   elements   that  

! combine   virtual   and   occupied   ground   state  !orbitals   in   the   excited   state   with  

!U ai = !Uia .  The  new  occupied   ! set   {" 'i ;i = 1, occ}  provides  us  with  the  excited  state  KS-­‐ determinant   ! ' = !1'!2' ...!i'! 'j ...!n'  from  which  we  can  express  the  corresponding  spin  

! and   space   density   matrices   s'  and   ! ' .   These   matrices   allow   us   subsequently   to   express   the   Kohn-­‐Sham   energy   of   ! '  as   E[ ! '+, ! '! ]  where   ! '+ = ! ' / 2 + s ' / 2  and  

! '! = ! ' / 2 ! s' / 2 .27     In   CV(2)-­‐DFT   the   summation   in   eq   1   is   carried   out   to   m=2     so   that   the   corresponding   orbitals   are   orthonormal   up   to   second   order   in   U.   From   that  

E[ ! '+, ! '! ]  is  expressed  to  second  order  in  U    from  the  second  order    density  and  spin   matrices   as   E[ ! (2)+, ! (2)! ] .   A   subsequent   optimization   of   E[ ! (2)+, ! (2)! ]  under   the   constraint27  that  exactly  one  electron  is  moved  from  the  density  space  spanned  by  

{" i ; i = 1, occ }  to   the   space   spanned   by   {" a ; a = 1, vir }  affords   after   applying   the   Tamm-­‐Dancoff  approximation28    the  eigenvalue  equation  

!

!                                                                       AU(I ) = ! (I )U (I )    (3)   where   Aai,bj = "ab"ij (# a $ # i ) + K ai,bj

(4).

!

 

ACS Paragon Plus Environment

5  

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 29

We shall refer to the CV(2)-DFT method in which we have used the Tamm-­‐Dancoff   approximation28    as simply CV(2)-­‐TD. C XC In eq 4 ! a and !i refer to ground state orbital energies. Further K ai,bj = K ai,bj + K ai,bj

where

K

C ai,bj

=

!

! 1 ! !a (1)!i (1) r !b (2)! j (2)dv1dv2 (5) 12

whereas XC(HF ) K ai,bj =!"

1

" ! (1)! (2) r a

i

!b (1)! j (2)dv1dv2 (6)

12

for Hartree Fock exchange correlation and XC(KS ) K ai,bj =

! ! ! (1)! (1) f (1)! (1)! (1)dv dv a

i

b

j

1

2

(7)

for KS exchange correlation. In eq 7 f is the energy kernel. We note that an expression similar to eq 3 can be derived4 from adiabatic TDDFT after application of the    Tamm-­‐ Dancoff  approximation.28    Thus  adiabatic  TDDFT-­‐TD  based  on  response  theory  and   CV(2)-­‐TD  based  on  variation  theory  are  mathematically  equivalent.    

After   solving   eq   3   we   apply   the   resulting   U to   eqs   1   and   2   in   an   all   order  

unitary   transformation   from   which   we   can   obtain   E[ ! (!)+, ! (!)" ]  corresponding   to   CV(∞)-­‐DFT.   In   general,   in   CV(n)-­‐DFT   we   use   the   U from   eq   3   to   express    

E[ ! (n)+, ! (n)! ] .  It  is  also  possible  to  determine   U directly  by  optimizing   E[ ! (n)+, ! (n)! ] .   In   that   case   we   talk   about   SCF-­‐CV(n)-­‐DFT.   27   We   shall   here   only   be   interested   in   CV(n)-­‐DFT   as   we   want   the   simplest   possible   theory   that   can   describe     π   →   π∗  

 

ACS Paragon Plus Environment

6  

Page 7 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

excitation   in   dyes.   However,   the   π   →   π∗   excitations   in   conjugated   hydrocarbons   have  been  studied  extensively  by    SCF-­‐CV(n)-­‐DFT.  27      

We   shall   now   turn   to   a   discussion   of   how   we   construct   the   proper   energy  

expression   for   excited   singlet   states   originating   from   a   closed   shell   ground   state.   Considering   first   a   spin-­‐conserving   transition   from   a   close   shell   ground   state   and   assume  without  loss  of  generality  that  the  transition  takes  place  in  the  α-manifold. In that case we can write the occupied excited state KS-orbitals generated from the unitary transformation of (1) as27

!i' = cos["# i ]!io$ + sin["# i ]!iv$ ; i = 1,occ / 2

(8)

and

!i' = !i ; i = occ / 2 + 1,occ ,

(9)

whereas the corresponding KS-determinant is given by

! M = "1'"2' ..."i'" 'j ..."n' .

(10)

Here ! M represents a mixed spin-state42 that is half singlet and half triplet. Further,

! i (i = 1,occ / 2) is a set of eigenvalues to (V!! )+ U!! W!! = 1" , (11) where ! is a diagonal matrix of dimension occ/2 whereas U!! is the part of the U matrix that runs over the occupied {!i ;i = 1,occ / 2} and virtual {!a ;i = 1,vir / 2} ground state orbitals of α-spin.27, 49 Further                                                                                   !io" =

occ/2

# (W

""

) ji ! j        (12)  

j

   and        

 

ACS Paragon Plus Environment

7  

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 29

vir/2

                                                                                                          ! = $ (V "" )ai# a      (13)   v" i

a

Finally ! is determined in such a way that

#

occ/2 i=1

sin 2 [!" i ] = 1 , corresponding to the

constraint that exactly one electron charge is involved in the transition. The orbitals defined in (12) and (13) have been referred to as Natural Transition Orbitals (NTO)50 since they give a more compact description of the excitations than the canonical orbitals. Thus, a transition that involves several i ! a replacements among canonical orbitals can often be described by a single replacement !io" # !iv" in terms of NTO’s. We note again that the set in (8) is obtained from the unitary transformation (1) among the occupied

{!i ;i = 1,occ / 2} and virtual {!a ;i = 1,vir / 2} ground state orbitals of α-spin with U represented by U!! . For a spin-flip transition from a closed shell ground state the unitary transformation (1) among the occupied ground state orbitals {!i ;i = 1,occ / 2} of α-spin and the virtual ground state orbitals {!a ;i = vir / 2 + 1,vir} of β-spin yields the occupied excited state orbitals

!i'' = cos["# i' ]!io$ + sin["# i' ]!i % ; i = 1,occ / 2 v

(14)

and

!i'' = !i ; i = occ / 2 + 1,occ ,

(15)

whereas the corresponding KS-determinant is given by

! T = "1'"2' ..."i'" 'j ..."n' .

(16)

Here ! T represents a triplet state. Further, ! i' (i = 1,occ / 2) are the eigenvalues to

(V !" )+ U !" W !" = 1# ' (17)

 

ACS Paragon Plus Environment

8  

Page 9 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

where U !" is the part of U that runs over the virtual ground state orbitals of β-spin and occupied ground state orbitals of α-spin. Finally                                                                                   !

o" i

=

occ/2

$ (W

#"

) ji ! j        (18)  

j=1

   and         a=vir

                                                                                                    !i " = v

%

(V "# )ai$ a .      (19)  

a=vir/2+1

With the energy of ! T given by ET and that of ! M as EM , we can write the singlet transition energy as42

!ES = 2EM " ET " E0 , (20) provided that U!! = U "! . This implies that V!! = V "! ; W !! = W "! and ! = ! ' . Further

!io" of (8) and (14) become identical as do the spatial parts of !iv" and !i " in the same two v

equations. In CV(∞)-DFT use is made of the U !" matrix obtained from a CV(2)-TD singlet calculation. Straightforward manipulations27 allow us finally to write down the singlet transition energy in a compact and closed form as !ES =

occ/2

$ sin ["# 2

i=1

+

i

](% i v& ' % i o& )

1 occ/2 occ/2 2 sin ["# j ]sin 2 ["# i ](K i o& i o& j o& j o& + K i v& i v& j v& j v& ' 2K i v& i v& j o& j o& + K i v( i v( j o& j o& ] (21) $ $ 2 i=1 j=1

occ/2 occ/2

$ $ sin["# i=1

j=1

j

]cos["# j ]sin["# i ]cos["# i ](2K i o& i v& j o& j v& ' K i o& i v( j o& j v( )

Here the indices i o! , j o! ,i v! and j v! refer to α-spin orbitals with the spatial parts

!io" ,! oj " ,!iv" and ! vj " , respectively. Likewise the indices i ! , j ! , i o

 

ACS Paragon Plus Environment

o

v!

and j ! refer to βv

9  

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

spin orbitals with the spatial parts !i " ,! j " ,!i " and ! j " , respectively. o

o

v

v

Page 10 of 29

In deriving

(21)

we have applied the Tamm-­‐Dancoff  approximation28    by  neglecting  integrals  of  the   type   K i o! i v! j v! j o! in  the  third  term  on  the  right  hand  side  of  (21)  as  well  as  numerically   very  small  contributions  that  are  third  order  in   sin[!" i ] .  These  approximations  are   introduced   to   ensure   that   (21)   for   excitations   that   are   described   by   several   displacements   converge   to   the   CV(2)-­‐TD   expression   for   a   singlet   transition,   as   we   shall   discuss   shortly.   The   expression   for   the   corresponding   triplet   transition   energy   reads     !ET =

occ/2

$ sin ["# 2

i=1

i

](% i v& ' % i o& )

1 occ/2 occ/2 2   + $ $ sin ["# j ]sin 2 ["# i ](K i o& i o& j o& j o& + K i v( i v( j v( j v( ' K i v( i v( j o& j o& ] (22) 2 i=1 j=1 occ/2 occ/2

$ $ sin["# i=1

j=1

j

]cos["# j ]sin["# i ]cos["# i ]K i o& i v( j o& j v(

We shall from now on refer to CV(∞)-DFT calculations in which (21) and (22) are used as CV(∞)-TD. Please note that for single displacement transitions where !" j = # / 2 for j=i and !" j = 0 for j ! i , we get the same energy whether the TD-approximation is used or not since only the first and second term in (22) contributes to !ES .

2.2    Computational  Details   All calculations have been performed with the developer’s version of the   Amsterdam   Density  Functional  (ADF)  package. The  Slater-­‐type  basis  supplied  by  ADF30  version   2010   was   of   triple-­‐ζ   quality31   with   an   additional   set   of   polarized   functions   (TZP).   From  extensive  TDDFT  studies23-­‐25  on  dyes,  we  estimate  that  our  excitation  energies  

 

ACS Paragon Plus Environment

10  

Page 11 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

are  within  0.1  eV  to  0.2  eV  of  those  obtained  by  a  converged  basis  set.  Use  was  made   of   the   local   density   approximation   LDA(VWN)32   in   the   parameterization   due   to   Vosko,   Wilk,   and   Nusair.   For   the   calculation   of   the   exact   Hartree-­‐Fock   exchange   integrals  over  Slater-­‐type  orbitals,  a  double  fitting  procedure  has  been  employed.33   All  ground  state  structures  were  adopted  from  the  work  of    Richard and Herbert16.   3  Results  and  Discussion   We   present   in   Table   1   the   calculated   1 A1g ! 1B2u and   1 A1g ! 1B3u  excitation   energies   !E( 1B2u )  and   !E( 1B3u ) ,respectively,   for   linear   acenes   employing   the   TDDFT  and  CV(2)-­‐TD  methods  in  conjunction  with  the  LDA(VWN)  functional.  Also   shown   are   the   calculated   oscillatory   strengths   of   the   two   excitations.   Extensive   studies16  have  shown  that  the   1 A1g ! 1B2u  transition  essentially  can  be  represented   by  the  HOMO  →  LUMO  promotion  whereas     1 A1g ! 1B3u  is  a  nearly  equal  mixture  of   the  HOMO-­‐1  →  LUMO  and  HOMO  →  LUMO+1    promotions.        Excitation  energies  ( !ES )  due  to  transitions  from  a  closed  shell  ground  state   to  a  singlet  excited  state  have  in  CV(2)-­‐TD  the  simple  form27   vir/2 occ/2

vir/2 occ/2

vir/2 occ/2

!ES = # # U aiU ai (! a " !i ) + 2 # # U aiU bj K ai,bj " # # U ai U bj K ai ,bj  .  (23)   a

i

a,b

i, j

a,b

i, j

Here   superscript   “bars”   refer   to   orbitals   of   ! ! spin   whereas   i, a without   bars   represent   ! ! orbitals.   The   corresponding   energies   in   TDDFT   has   a  somewhat  more   complex  form4  but  depend  again  on   (! a ! !i ) ,     K ai,bj  ,   K ai ,bj  and     U .  

 

ACS Paragon Plus Environment

11  

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 29

It   follows   from   Table   1   and   Figure   1a   that   the   experimental   energy   gap   !E = !E( 1B2u ) ! "E( 1B3u )  starts   out   positive   at   naphthalene   (nr=2)   with   Δ! =

0.53  !"  before   turning   negative   at   anthracene   (nr=3)   where   Δ! = −0.04  !".  For   larger   linear   acenes    Δ!  becomes   increasingly   negative   reaching  Δ! = −0.85  !"  at   hexacene.   Thus,   experimentally,  Δ!( !!!! )  is   seen   to   drop   faster   in   energy   than   Δ!( !!!! ).     The   ordering   of   the   calculated   CV(2)-­‐TD   excitation   energies   based   on   LDA(VWN)  is  correct  for  naphthalene  as  well  as  for  all  other  linear  acenes,  however,   the   gap   differs   from   the   experimental   Δ!  by   almost   0.4   eV   for   naphthalene,   anthracene  and  hexacene.  This  difference  is  smaller  for  napthacene  and  pentacene,   Table   1   and   Figure   1b.   It   can   be   seen   that   the   main   contribution   to   this   deviation   comes  from  the  underestimation  of  Δ!( !!!! ).  Thus,  the  root  mean  square  deviation   (RMSD)  value  for  Δ!( !!!! )  is  0.49  eV,  while  it  is  only  0.13  eV  for  Δ!( !!!! ).   As   for   the   TDDFT   results   based   on   LDA(VWN),   the   deviation   from   the   experimental  gaps  is  larger  in  absolute  terms  and  the  calculated  Δ!  has  the  wrong   sign   for   naphthalene,   Table   1   and   Figure   1c.   It   happens,   because   the  Δ!( !!!! )   values  for  TDDFT    are  lower  than  those  for  CV(2)-­‐TD    while  the    Δ!( !!!! )  estimates   for  TDDFT  are  higher  than  for  CV(2)-­‐TD.  As  a  result,  the  RMSD  value  for  Δ!( !!!! )   increases   for   TDDFT   by   approximately   0.2   eV   and   reaches   0.71   eV.     On   the   other   hand,   on   average   the   TDDFT   estimate   of  Δ!( !!!! )  is   as   accurate   as   for   CV(2)-­‐TD.   Thus,   the   TDDFT  Δ!( !!!! )  values   for   napthacene,   pentacene   and   hexacene   are   closer   to   experiment   than   those   due   to   CV(2)-­‐TD   while   the   opposite   is   true   for  

 

ACS Paragon Plus Environment

12  

Page 13 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

naphthalene  and  anthracene.  The  RMSD  value  for  TDDFT  is  0.14  eV  compare  to  0.13   eV  for  CV(2)-­‐TD,  Table  1.    

It  follows  from  the  discussion  given  above  that  neither  CV(2)-­‐TD  nor  TDDFT  

are  able  to  give  a  quantitative  description  of    Δ!  as  a  function  of  nr  with    LDA,    in  line   with  previous  TDDFT  studies2,3,8,9    using  both  pure  density  functionals  and  hybrids.   The   source   of   the   error   is   in   all   cases   primarily  Δ!( !!!! )  that     is   too   low   compared   to  experiment.  However,  Δ!( !!!! )  is  in  addition  seen  to  be  slightly  too  high.       We  have  in  order  to  analyze  our  findings  further  displayed  the  orbital  energy   dependent  part                                                            Δ! =

!"# !

!

!"" !

!

!!" !!" (!! − !! )    (24)  

in   the   expression   for    Δ!!  according   to   CV(2)-­‐TD   ,   eq   23,   along   with   the   K-­‐integral   dependent  part    Δ!! − Δ!  in  Table  2  and  Figures  2a-­‐b  as  a  function  of  nr  for  LDA  in   the  case  of    Δ!( !!!! )  and  Δ!( !!!! ).    We  see  that  both  the  slope  and  the  absolute   value  of      Δ!( !!!! )  and  Δ!( !!!! )  are  determined  by  Δ!  and  that  the  contributions   to   Δ!!  from   the   K-­‐integrals   amount   to   0.3   eV   for     Δ!( !!!! )  and   0.1   eV   for     Δ!  

!

!!!  ,where  both  contributions  diminish  slightly  with  nr.     Turning   now   to   the   CV(∞)-­‐DFT   scheme   we   note     that   the   excitation   energy  

(21)  for  a  transition  that  involves  a  single  promotion  ! → !  such    as   1 A1g ! 1B2u     has   the  simple  form27                                      Δ!!!"

!

!

!

! → ! = !! − !! + ! !!!!! + ! !!!!! − 2!!!"" + !!!!!    (25).  

Here  i  is  the  HOMO      and  a  the  LUMO.  For  CV(2)  we  obtain  for  the  same  transition   according  to  eq    23  

 

ACS Paragon Plus Environment

13  

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

                               Δ!!!"

!

Page 14 of 29

! → ! = !! − !! + 2!!"!" − !!!!!      (26).  

For   HF   these   two   expressions   are   identical27   since  !!!!! = !!!!! = 0  and    !!"!" = −!!!"" .  However  for  any  of  the  popular  functionals  this  is  not  the  case.  Thus,  the  two   expressions  will  give  rise  to  different  excitation  energies  for  the  same  functional,  as   we   shall   see.     For   transitions   consisting   of   two   or   more   promotions   such   as   1

!" !

A1g ! 1B3u ,   the   expression   of   eq   (21)   for  Δ!!

 contains   contributions   that   are  

second   and   fourth   order   in   sin.   The   contributions   from   the   latter   terms   diminish   rapidly   in   magnitude   with   the   number   of   promotions   due   to   the   constraint  

#

occ/2 i=1

sin 2 [!" i ] = 1 .   Thus,   for   four   equally   participating   promotions   with  

sin 2 [!" i ] = 1 / 4  (i=1,4),   the   fourth   order   prefactors   sin 4 [!" i ] are   small     (1/16)   and   the  terms  in  (21)  to  second  order  in  sin  will  dominate  since   sin 2 [!" i ] = 1 / 4  (i=1,4)  .   In  the  limit  of  many  more  participating  promotions  (21)  will  converge  to  (23).      

It  follows  from  Table  3  and  Figure  1d  that  the  CV(∞)-­‐DFT  results  using  LDA  

are   in   excellent   agreement   with   experiment   for   both  Δ!( !!!! )  and   Δ!( !!!! )   throughout  the  range  of  linear  acenes.  The  RMSD  for  Δ!( !!!! )  is  0.06  eV  whereas   that  for  Δ!( !!!! )  is  0.13  eV.    Thus,  CV(∞)-­‐DFT  clearly  represents  an  improvement   over   TDDFT   and   CV(2)-­‐TD   for   LDA.   The   improvement   is   as   anticipated   most   noticeable   for  Δ!( !!!! )  where   the   RMSD   was   0.71   eV   for   TDDFT   and   0.49   eV   for   CV(2)-­‐TD.  For  Δ!( !!!! )  all  three  methods  have  a  similar  RMSD.      

The magnitude of 1/2!!!!! + 1/2!!!!! − 2!!!"" + !!!!! from eq 25 is given in

Table 4 and Figure 2c for !!!! as Δ!( !!!! )-­‐  Δ!

!

!!! . It is on the average 0.8 eV and

is seen to decrease slightly with n. The magnitude of this term depends strongly on the  

ACS Paragon Plus Environment

14  

Page 15 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

spatial extents of a and i. If the densities of a and i are strongly overlapping the first two terms might be canceled by the last two. This has been found for some π   →   π∗   excitation.27   At   the   other   extreme   is   charge   transfer   transitions35-­‐38   where   there   is   no   overlap   and   where   the   centers   of   the   two   densities   are   separated     by   a   large   distance   R.   In   that   case   the   two   last   terms   reduce   to   1/R.     Many   studies have   shown   that  charge  transfer  transitions  are  better  described  by  (25)  than  by  (26).35-­‐39    In  the   case   at   hand   the   0.8   eV   supplied   by   1/2!!!!! + 1/2!!!!! − 2!!!"" + !!!!! is just enough to bring the CV(∞)-DFT values calculated for Δ!( !!!! ) in line with experiment whereas the corresponding CV(2)-TD contribution given by 2!!"!" − !!!!!   is too small (0.30 eV). Given that the excitation energy for one-orbital transitions in CV(∞)-DFT is expressed by (10) rather than (11), we can in general expect it to differ from that calculated by CV(2)-TD or TDDFT irrespective of whether or not it exhibits charge transfer character. Also shown in Table 4 and Figure 2d is Δ!( !!!! )-­‐  Δ!

!

!!!  calculated by

CV(∞)-DFT for the linear acenes. It is much smaller than Δ!( !!!! )-­‐  Δ! more in line with the corresponding Δ!( !!!! )-­‐  Δ!

!

!

!!!

and

!!! values obtained by CV(2)-TD

or TDDFT, Table 2 and Figure 2b. This is consistent with that 1Lb  (1B3u)  is  described  by   more  than  one  orbital  substitution  and  the  fact  that  the  energy  expression  (21)  for   CV(∞)-TD in the limit of excitations with several orbital replacements affords CV(2)-TD transition energies given by (23). Kuritz19  et   al.  have  provided  an  interesting  analysis   of  how  LRC  functionals  improve  the  description  of  the  1La  transitions.  

 

ACS Paragon Plus Environment

15  

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 29

We have extended our benchmark calculations also to include the 15 nonlinear acenes shown in Figure 3. The resulting singlet transition energies involving 1La and 1Lb are shown in Table 5 for TDDFT, CV(2)-TD and CV(∞)-TD based on LDA. For 1La CV(∞)-TD with a RMSD of 0.24 eV is seen to perform better than

CV(2)-TD

(RMSD=0.40 eV) and especially TDDFT ( 0.52 eV). In fact our results are of a similar quality to the best results obtained by long-­‐range   corrected   (LRC)   functionals.16     We   note   again   that   the   1La   transitions     involves   a   single   HOMO   !  LUMO   orbital   displacement   with     !" i = # / 2 .   For   1Lb   all   three   methods   performs   equally   well   with   RMSD’s   of   0.16   eV   (TDDFT),   0.15   eV(CV(2)-­‐TD)   and   0.19   eV   (CV(∞)-­‐TD),   respectively.  It  is  interesting  to  note  that  the  LRC-­‐functionals16    in  this  case  performs   much  poorer  with  RMSD’s  around  0.4  eV.    Thus  ,  CV(∞)-­‐TD  at  the  simple  LDA  level   is  the  only  scheme  of  the  methods  discussed  here  that  gives  a  balanced  description   of  π  →  π∗    transitions  involving  a  single  orbital  displacement  (1La)  as  well  as  a  π   →  π∗    transitions  with  more  than  one  displacement.    

 

 

3.  Concluding  Remarks    

We   have   here   tested   the   all   order   constricted   variational   density   functional  

theory  CV(∞)-­‐TD  in  conjunction  with  LDA  as  a  possible  inexpensive  method  for  the   computational  screening  of  organic  dyes.  Our  initial  test  set  was  made  up  of  linear   poly-­‐acenes   ranging   from   naphthalene   with   two   to   hexacene   with   six   rings.   For   these  systems  we  studied  the  energy  difference  Δ!  between  the  excitation  energies   of   the   first   two   π   →   π∗     transitions   1La   (1B2u)   and   1Lb   (1B3u)   as   a   function   of   the  

 

ACS Paragon Plus Environment

16  

Page 17 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

number   of   rings   nr.     These   systems   are   known   to   be   challenging   for   DFT   based   electronic  structure  methods.      

Our   study   revealed   that   CV(∞)-­‐TD   combined   with   LDA   gave   as   accurate   a  

description   of     Δ!  as   the   sophisticated   DFT/MRCI11-­‐12   and   TDDFT/LRC  

13-­‐16  

approaches   for   linear   acenes   while   representing   a   clear   improvement   over   TDDFT2,3,8,9  using  pure  functionals  or  hybrids.  This  is  interesting  given  the  low  cost   and  simplicity  of  the  LDA  functional  and  the  computational  efficiency  of    CV(∞)-­‐TD   which  is  a  factor  two  more  expensive  than    TDDFT  or  CV(2)-­‐TD  schemes  alone.    

It  is  very  well  known  that  transitions  involving  a  single  substitution  such  as  

1L   (1B )   a 2u

often   are   better   described   by   (25)   that   originates   from   the  ∆!"#  39-­‐46  

energy  expression  than  from  (26)  due  to  TDDFT-­‐TD  .  This  point  has  recently  been   demonstrated   also   to   be   the   case   for   π   →   π∗     transitions27,47     in   conjugated   hydrocarbons.   Also   Casida48   has   suggested   to   use   (25)   in   place   of   (26)   for   single   substitution  transitions  involving  charge  transfer  as  have  others.35-­‐37  Thus,  as  such   our  CV(∞)-­‐TD  method  does  not  bring  anything  new.    

The   novel   aspect   of   our   scheme   is   that   it   provides   a   seamless  

transition   from   a   single   substitution   excitation   where   the   ∆!"#  type   energy   expression   (25)   is   required   to   multi-­‐replacement   transitions   where   (26)   is   appropriate.   The   all   order   constricted   variation   theory   represents   in   addition   a   sound  theoretical  framework  that  contains  both    ∆!"#  and  adiabatic  TDDFT-­‐TD  as   special  cases.  We  should  finally  note  that  that  the  CV(∞)-­‐TD  scheme  is  a  simple  form   of     more   refined   theories   such   as   SCF-­‐CV(∞)-­‐DFT27   in   which   U is   optimized   with  

 

ACS Paragon Plus Environment

17  

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 29

respect  to   E[ ! (n)+, ! (n)! ]  rather  than   E[ ! (2)+, ! (2)! ]  and  R-­‐CV(∞)-­‐DFT    or  RSCF-­‐CV(∞)-­‐ DFT  in  which  we  also  introduce  relaxation  of  the  orbitals  not  participating  directly   in   the   excitation.29   However,   in   spite   of   its   simplicity   the   CV(∞)-­‐TD     scheme   was   shown   recently   to   afford   an   average   deviation   from   experiment   of     approximately   0.3  eV    for  both  semi-­‐local  and  hybrid  functionals  in  an  extensive  study  of  π  →  π∗     transitions27,47     involving     conjugated   hydrocarbons.   The   current   study   on   linear   acenes  adds  further  to  the  notion  that  CV(∞)-­‐TD    might  be  a  useful  and  cost  efficient   supplement   to   TDDFT   in   the   study   of   dyes.   We   have   in   addition   provided   an   extensive  theoretical  analysis  of  how  CV(∞)-­‐TD  complement  TDDFT.     Acknowledgement.   This   work   was   supported   by   NSERC.   The   computational   resources   of   WESTGRID   were   used   for   all   calculations.   T.Z.   thanks   the   Canadian   Government   for   a   Canada   Research   Chair   and   the   Humboldt   foundation   for   a   generous  award.    

 

ACS Paragon Plus Environment

18  

Page 19 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

References (1) Platt, J. R. J. Chem. Phys. 1949, 17, 484–496. (2) Grimme, S.; Parac, M. Chem. Phys. Chem. 2003, 4, 292–295. (3) Parac, M.; Grimme, S. Chem. Phys. 2003, 292, 11–21. (4) Casida, M. E. In Recent Advances in Density Functional Methods; D. P. Chong, Ed.; World Scientific: Singapore, 1995; Vol. 1, pp. 155-192. (5) Gross, E. K. U.; Dobson, J. F.; Petersilka, M. Top. Curr. Chem. 1996, 181, 81–172. (6) Bauernschmitt, R.; Ahlrichs, R. Chem. Phys. Lett. 1996, 256, 454–464. (7) Time-Dependent Density Functional Theory, Lecture Notes Physics 706; Marques, M. A. L., Ulrich, C. A., Nagueira, F., Rubio, A., Burke, K., Gross, E. K. U., Eds.; Springer: Berlin-Heidelberg, 2006. (8) Jacquemin, D.; Wathelet, V.; Perpete, E. A.; Adamo, C. J. Chem. Theory Comput. 2009, 5, 2420–2435. (9) Goerigk, L.; Grimme, S. J. Chem. Phys. 2010, 132, 184103. (10) Dierksen, M.; Grimme, S. J. Phys. Chem. A 2004, 108, 10225-10237. (11) Grimme, S.; Waletzke, M. J. Chem. Phys. 1999, 111, 5645–5655. (12) Marian, C.; Gilka, N. J. Chem. Theory Comput. 2008, 4, 1501–1515. (13) Jacquemin, D.; Perpete, E. A.; Scuseria, G. E.; Ciofni, I.; Adamo, C. J. Chem. Theory Comput. 2008, 4, 123–135. (14) Jacquemin, D.; Perpete, E. A.; Ciofni, I.; Adamo, C. Theor. Chem. Acc. 2011, 128, 127–136. (15) Wong, B. M.; Hsieh, T. H. J. Chem. Theory Comput. 2010, 6, 3704–3712. (16) Richard, R. M.; Herbert, J. M. J. Chem. Theory Comput. 2011, 7, 1296–1306. (17) Stein, T.; Kronik, L.; Baer, R. J. Am. Chem. Soc. 2009, 131, 2818–2820. (18) Stein, T.; Kronik, L.; Baer, R. J. Chem. Phys. 2009, 131, 244119. (19) Kuritz, N.; Stein, T.; Kronik, R. B. L. J. Chem. Theory Comput. 2011, 7, 2408–2415. (20) Grimme, S.; Neese, F. J. Chem. Phys. 2007, 127, 154116. (21) Grimme, S. J. Chem. Phys. 2006, 124, 034108. (22) Karton, A.; Tarnopolsky, A.; Lamere, J. F.; Schatz, G. C.; Martin, J. M. L. J. Phys. Chem. A 2008, 112, 12868–12886. (23) Goerigk, L.; Moellmann, J.; Grimme, S. Phys. Chem. Chem. Phys. 2009, 11, 4611– 4620. (24) Goerigk, L.; Grimme, S. J. Phys. Chem. A 2009, 113, 767–776. (25) Goerigk, L.; Grimme, S. J.Chem.Theory Comput. 2011,7,3272-3277. (26) Head-Gordon, M.; Rico, R. J.; Oumi, M.; Lee, T. J. Chem. Phys. Lett. 1994, 219, 21–29. (27)  Ziegler,T.;  Krykunov,  M.  and    Cullen,  J.    J.  Chem.  Phys.  2012,  136,  124107.     (28)  Hirata,  S;  Head-­‐Gordon  M.  Chem.  Phys.  Lett.  1999,  314,  291-­‐299.   (29) Ziegler, T; Krykunov, M. J. Chem. Phys. 2010, 133, 074104. (30)   te   Velde,   G.;   Bickelhaupt,   F.   M.;   van   Gisbergen,   S.   J.   A.;   Fonseca   Guerra,   C.;   Baerends,  E.  J.;  Snijders,  J.  G.;  Ziegler,  T.  J.  Comput.  Chem.  2001,  22,  931-­‐967.   (31)  Van  Lenthe,  E.;  Baerends,  E.  J.  J.  Comput.  Chem.  2003,  24,  1142-­‐1156.   (32)  Vosko,  S.  H.;  Wilk,  L.;  Nusair,  M.  Can.  J.  Phys.  1980,  58,  1200-­‐1211.  

 

ACS Paragon Plus Environment

19  

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 29

(33)   Krykunov,   M.   ;   Ziegler,   T.     and     van   Lenthe,   E.     J.   Phys.   Chem.   A,   2009,   113,   11495–11500   (34)  Biermann,B.;  Schmidt,W.    J.Am.Chem.Soc.  1980,102,3163-­‐3173   (35) Ziegler, T.; Seth, M.; Krykunov, M.; Autschbach, J.; Wang, F. (Theochem) J.Mol. Struct. 2009, 914, 106–109. (36) Ziegler, T.; Seth, M.; Krykunov, M.; Autschbach, J. J. Chem. Phys. 2008, 129, 184114. (37) Ziegler, T.; Seth, M.; Krykunov, M.; Autschbach, J.; Wang, F. J.Chem. Phys. 2009, 130, 154102. (38)  Cullen,J.;  Krykunov,M.  and  Ziegler,T.  Chem.Phys.  2011,391,11-­‐18.   (39)  Ziegler,T.;  Rauk,  A.  Baerends,  E.J.    Chem.  Phys.,  1976,  16,  209-­‐217.     (40)  Slater,  J.C.  and  Wood,J.H.    Int.J.  Quantum  Chem.  1971,  Suppl.  4  ,  3-­‐34.   (41)  Slater,J.C.  Adv.  Quantum.Chem.  1972,  6  ,1-­‐31.   (42)  Ziegler,T.;  Rauk,  A.    and  Baerends,  E.J.    Theor.Chim.  Acta  1977,  43,  261-­‐271.   (43  )  Levy,M.  and  Perdew,J.P.    Phys.Rev.  1985,  A  32,  2010-­‐2021.   (44)    Nagy,A.  Phys.  Rev.  A  1996,  53,  3660-­‐3663.   (45)  Besley,N;    Gilbert,A    and    Gill,  P.    J.  Chem.  Phys.  2009,130,  124308.    (46)  Gavnholt,  J.  ;  Olsen,  T.;    Engelund,  M;    Schiotz,J.      Phys.  Rev.  B    2008,  78,  075441.   (47)  Kowalczyk,T.;  Yost,S.R.;  Van  Voorhis,  T.    J.Chem.Phys.  2011,  134,054128.   (48)  Cassida,  M.E.  ;  Gutierrez,  F.  ;  Guan,  J.  Gadea,  F.-­‐X.  ;  Salahub,  D.    and  Daudey,  J.-­‐P.     J.  Chem.  Phys.  2000,113,  7062-­‐7071.   (49)  Amos,  A.T;  Hall,  G.G.  Proc.  Roy.Soc.  1961,    A263  ,  483-­‐493.     (50)  Martin,R.L.      J.Chem.Phys.  2003,  118  ,  4775-­‐4777.              

 

ACS Paragon Plus Environment

20  

Page 21 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

  Table 1. LDA(VWN) based 1La(1B2u) and 1Lb(1B3u) singlet excitation energies (in eV) for acenes with CV(2)-TD and TDDFT c b Experiment CV(2)-TDd f TDDFT    

 

1B 2u

1B

a

 

a 1B 1B nr ∆E ∆E ∆E 3u 2u 3u 2 4.66 4.13 -0.53 4.27 4.21 0.06 0.0545 0.0000 4.07 4.41 0.34 3 3.60 3.64 -0.04 3.13 3.60 -0.47 0.0578 0.0003 2.91 3.78 -0.87 4 2.88 3.39 -0.51 2.38 3.21 -0.83 0.0533 0.0013 2.15 3.35 -1.20 5 2.37 3.12 -0.75 1.86 2.94 -1.08 0.0469 0.0033 1.62 3.08 -1.46 6 2.02 2.87 -0.85 1.47 2.74 -1.27 0.0399 0.0071 1.22 2.86 -1.64 RMSD 0.49 0.13 0.71 0.14 a ∆E = ∆E(1B2u) − ∆E(1B3u). bOscillator strength.cRef. 34.dTamm-Dancoff approximation. eNumber  of  rings e

1B 2u

1B 3u

a

1B 2u

ACS Paragon Plus Environment

1B 3

21  

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 29

  Table  2.  Contributions  from    orbital  energies  and  K   integrals  to  the  calculated  excitation  energies  (eV)   for  CV(2)-­‐TDd  and  LDA.     Cal. 1B2u Cal. 1B3u   a                        ! cn   ∆!          Δ!! − Δϵ   r ∆! (Δ!! − Δϵ) 2   3.95   0.32   4.14   0.07   3   2.83   0.30   3.55   0.05   4   2.11   0.27   3.16   0.05   5   1.60   0.26   2.90   0.04   6   1.23   0.24   2.71   0.03   aOrbital  energy  contributions  to  the  excitation  energies,     see  eqs  25  and  26.     bContributions  from  K-­‐integrals  to  the  excitation  energies.   cNumber  of  rings.   d Tamm-Dancoff approximation.      

 

 

ACS Paragon Plus Environment

22  

Page 23 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

    Table 3. CV(∞)-TD singlet excitation energies (in eV) for linear acenes with LDA. Number bExperiment LDA a 1 1 of rings B2u B3u ∆E 1B2u 1B3u a∆E 2 4.66 4.13 0.53 4.73 4.39 0.34 3 3.60 3.64 -0.04 3.68 3.73 -0.05 4 2.88 3.39 -0.51 2.91 3.32 -0.41 5 2.37 3.12 -0.75 2.35 3.03 -0.68 6 2.02 2.87 -0.85 1.93 2.82 -0.89 RMSD 0.06 0.13 a ∆E = ∆E(1B2u) − ∆E(1B3u). bRef. 34      

 

 

ACS Paragon Plus Environment

23  

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 29

    Table  4.  Contributions  from    orbital  energies  and  K   integrals  to  the  calculated  excitation  energies  (eV)  using   CV(∞)-­‐TD  and  LDA.     Cal. 1B2u Cal. 1B3u   a                        ! cn   ∆!          Δ!! − Δϵ   r ∆! (Δ!! − Δϵ) 2   3.95   0.78   4.14   0.25   3   2.83   0.85   3.55   0.18   4   2.11   0.80   3.16   0.16   5   1.60   0.75   2.90   0.13   6   1.23   0.70   2.71   0.11   aOrbital  energy  contributions  to  the  excitation  energies,     see  eqs  25  and  26.     bContributions  from  K-­‐integrals  to  the  excitation  energies.     cNumber  of  rings.          

 

ACS Paragon Plus Environment

24  

Page 25 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

Table 5. LDA(VWN) based 1La(1B2u) and 1Lb(1B3u) singlet excitation energies (in eV) for nonlinear acenes with TDDFT , CV(2)-TD and CV(∞)-TD 1 1 La Lb Moleculea

TDDFT CV(2)b CV(∞)

b

Expc TDDFT CV(2)b CV(∞)b Exp.c

1

3.88

3.99

4.09

4.24

3.59

3.61

3.74

3.76

2

3.39

3.60

3.69

3.71

3.46

3.47

3.58

3.53

3

4.16

4.27

4.41

4.54

3.74

3.76

4.00

3.89

4

3.37

3.51

3.83

3.89

3.43

3.41

3.58

3.43

5

2.96

3.16

3.98

3.63

3.10

3.07

3.68

3.22

6

2.53

2.78

2.90

2.85

3.22

3.23

3.37

--

7

3.36

3.51

3.61

3.73

3.22

3.22

3.33

3.37

8

2.84

3.02

3.56

3.22

3.07

3.09

3.43

3.06

9

3.29

3.35

3.55

3.80

3.14

3.14

3.26

3.30

10

3.07

3.18

3.37

3.84

2.98

2.99

3.08

3.32

11

3.08

3.18

4.12

3.84

3.20

3.21

3.31

3.31

12

3.11

3.21

3.47

3.86

3.01

3.03

3.17

3.14

13

2.72

2.84

3.34

3.40

3.02

3.02

3.22

3.23

14

2.55

2.80

3.02

2.86

2.99

2.99

3.08

--

15

3.69

3.90

3.87

3.72

3.85

3.86

3.94

--

RMSD

0.52

0.40

0.24

0.16

0.15

0.19

a

The structures are shown in Fig. 1. Tamm-Dancoff approximation. c The experimental values are taken from Ref. 2.     b

 

 

ACS Paragon Plus Environment

25  

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1. Excitation energies for the 1La and 1Lb transitions in linear acenes as a function of the number of rings according to (a)experiment, (b) CV(2)-TD, (c) TDDFT and (d) CV(∞)-TD. 153x107mm (150 x 150 DPI)

ACS Paragon Plus Environment

Page 26 of 29

Page 27 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

Figure 2. Contributions from orbital energies and K integrals to the calculated excitation energies (eV) in linear acenes for CV(2)-TD with respect to (a) 1B2u (1La) and (b) 1B3u (1Lb) as well as CV(∞)-TD with respect to (c) 1B2u (1La) and (d) 1B3u (1Lb) 150x107mm (150 x 150 DPI)

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 3. Structures for nonlinear acenes 183x238mm (150 x 150 DPI)

ACS Paragon Plus Environment

Page 28 of 29

Page 29 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

graphical abstract 215x279mm (150 x 150 DPI)

ACS Paragon Plus Environment